Pb-Based Ferroelectric Nanomaterials
 1588830640

  • 0 0 0
  • Like this paper and download? You can publish your own PDF file online for free in a few minutes! Sign Up
File loading please wait...
Citation preview

Encyclopedia of Nanoscience and Nanotechnology

www.aspbs.com/enn

Pb-Based Ferroelectric Nanomaterials Ki Hyun Yoon Yonsei University, Seoul, Korea

Dong Heon Kang University of Suwon, Suwon, Korea

CONTENTS 1. Introduction 2. Synthesis of Lead-Based Ferroelectric Thin Films 3. Physical Properties of PMN and PZT Thin Films 4. Electrical Fatigue of PZSTN Thin Films 5. Conclusions Glossary References

1. INTRODUCTION There has been significant interest in the ferroelectric and antiferroelectric perovskite thin films such as Pb(Mg1/3 · Nb2/3 O3 (PMN), Pb(Zr,Ti)O3 (PZT), (Ba,Sr) TiO3 (BST), and Pb(Zr,Sn,Ti,Nb)O3 (PZSTN) because of their potential advantages for electro-optic and microelectronic applications including capacitors, pyroelectric detectors, ultra-highdensity data storage systems, nonvolatile semiconductor memories, piezoelectric sensors and devices [1–9]. Recently, with miniaturization and integration of electronic circuits, the electronic devices with high performance in their small size up to nano scale are increasingly required. Current microelectromechanical systems (MEMS) technology, which is typically constructed on the micrometer scale, is being pushed into the nanometer range as well, leading to the nanoelectrochemical system (NEMS) [10–13]. It is well known that such a regime will be attainable by not only nanomachining technique such as lithography and printing but also proper thin-film techniques with homogeneous and fine grains under 100 nm in size. However, in the early stage, the interest in this nanotechnology has been mainly focused on the ceramic powders, which increased enormously with the availability and the possibility of consolidation into dense nanostructured ceramics. The most ISBN: 1-58883-064-0/$35.00 Copyright © 2004 by American Scientific Publishers All rights of reproduction in any form reserved.

important factors limiting this goal are powder agglomeration, contamination during processing, and the residual pores on various length scales in the bodies [14–16]. Nowadays, the nanocrystalline structure has been obtained more frequently and actually through the thin-film process rather than through the powder process. For such submicron-scale applications very fine-grained thin films with highly uniform and highly textured microstructures are preferable. Thus intensive and extensive studies have been attempted to prepare the applicable thin films with nanocrystalline structures by using various fabrication methods, such as wet-chemical processes, physical and chemical vapor depositions [17–31]. In the case of solution-based (wet)-chemical method, it has been known to offer numerous advantages, including excellent compositional control, uniform homogeneity, ease of fabrication over large area, and especially low cost. And so several wet-chemical methods utilizing metallorganic compounds and sol–gel processing have been extensively employed to prepare thin films of the lead-based ferroelectric materials [32, 33]. The methoxyethanol route in which 2-methoxyethanol was conventionally used as a solvent has been applied to the alkoxide-based sol–gel process for the film preparation because 2-methoxyethanol can act as a bridging ligand and has a sufficient solubility for alkoxides and salts [34, 35]. However, lead-based thin films deposited on the Pt or oxide electrode by conventional methoxyethanol sol–gel process suffer from the presence of irregular microstructures which are detrimental to submicron device performance. Modified processes using different solvents and/or chelating agents are often applied to overcome some problems of solution and film such as the instability against moisture and difficulty of multicomponent synthesis [36–38]. In this study we introduced the solution-derived lead-based perovskite thin films with different chemical compositions and electric phases such as ferroelectric Pb(Mg1/3 Nb2/3 O3 (PMN), ferroelectric PbTiO3 (PT), antiferroelectric PbZrO3 (PZ), ferroelectric/antiferroelectric Pb(Zr1−x Tix O3 (PZT), and ferroelectric/antiferroelectric Encyclopedia of Nanoscience and Nanotechnology Edited by H. S. Nalwa Volume 8: Pages (435–444)

436 Pb099 [(Zr06 Sn04 1−X TiX ]098 Nb002 O3 (PZSTN) by using the modified and conventional methoxyethanol routes [39–41]. Their microstructures could be controlled in the respect of solution chemistry, phase transformation kinetics, and composition control, respectively. The solution processing, film heat-treatment condition, film composition, and deposition sequence are widely discussed in terms of the nanocrystallization of the thin film. Considering the fact that in case of chemical solution-derived thin films, most of fabrication processes were limited to the hydrothermal and spray pyrolysis methods [19–27], this study seems to be meaningful. Furthermore, the recent approaches have been to acquire a long-term reliability of the ferroelectric thin films including resistance to electric fatigue, imprint, and leakage. It has been suggested that the fatigue in the antiferroelectric ceramics and thin films is much less severe than that in the ferroelectric counterparts [42, 43]. Also, the modification of nanocrystalline size is thought to be related closely with electrical fatigue. So the novel method has been introduced to modify nanocrystalline microstructure and to study its effect on electrical fatigue in PZT and PZSTN with the Pt electrode system. Therefore, the main topics of this chapter are to explain and analyze the effects of the modified solution chemistry, crystallization, and compositional variation on the nano-scale microstructure and related characteristics of the ferroelectric and antiferroelectric lead-based thin films.

2. SYNTHESIS OF LEAD-BASED FERROELECTRIC THIN FILMS

Pb-Based Ferroelectric Nanomaterials

B-site precursors were mixed and allowed to react at room temperature for 3 h to form the PMN solution. For the preparation of PT and PZ solutions, titanium isopropoxide and AcAc (R = 1; where R = mole of AcAc/mole of metal alkoxides) were added in 2-methoxyethanol. The refluxing and distillation of this mixture resulted in goldcolored solutions with a concentration of about 0.5 M. After refluxing for 3 h in air, the solution was cooled to room temperature. Finally, the PT chemical solution was prepared by mixing of the Pb and Ti solutions and diluting with 2-methoxyethanol. Similarly, PZ solution was prepared by the same solution processing method mentioned above. Pb(Zr05 Ti05 O3 (PZT) solution was obtained by mixing and reacting equal amounts of PT and PZ precursors. For (x)PMN-(1 − x)PZT and (x)PMN-(1 − x)PT solutions, the corresponding amounts of PZT, PT, and PMN precursors were mixed and allowed to react for 3 h to obtain a desired composition. In the conventional process, all starting sources of lead for A-site cation and others such as Zr, Nb, Sn, Ti for B-site cation were refluxed and distilled in 2-methoxyethanol as described elsewhere [42, 43]. Then each solution was admixed to the corresponding Pb099 [(Zr06 Sn04 1−X TiX ]098 Nb002 O3 (PZSTN), PZ, and PZT compositions. Various PZSTN compositions with different electric phases were obtained by changing x; as x = 003 for antiferroelectrics (40/3/2) and x = 015 for ferroelectrics (40/15/2). All mixed-precursor solutions prepared by the modified and conventional processes were partially hydrolyzed to give a 0.3 M stable sol by diluting them with 2-methoxyethanol. After refluxing at room temperature for 24 h in air, they were aged for 24 h for the film coating process.

2.1. Solution Preparation The Pb-based complex perovskite films were prepared by a conventional methoxyethanol chemical solution deposition [44] and a modified version of the route described by Yi et al. [35] and Park et al. [39, 45]. The starting materials were lead acetate (Pb(CH3 COO2 3H2 O), zirconium n-propoxide (Zr(O-nC3 H7 4 , magnesium ethoxide (Mg(OC2 H5 2 , tin acetate (Sn(CH3 COO)4 , titanium isopropoxide (Ti(O-iC3 H7 4 , and niobium ethoxide (Nb(OC2 H5 2 . Propylenglycol, deionized water, and 2-methoxyethanol (CH3 OC2 H4 OH) were used as a solvent. For the PMN and PZT precursors prepared by modified process, triethanolamine (TEA) and acetylacetone (AcAc) were used as a complexing agent to control hydrolysis reaction with metal alkoxides, respectively. In the case of PMN solution, the A-site and B-site precursors of ABO3 perovskite structure were prepared separately. The appropriate amounts of the B-site precursors, such as magnesium ethoxide and niobium ethoxide, were mixed in 2-methoxyethanol and heated to ∼110  C for 24 h in a dry atmosphere. The solutions were then cooled to room temperature and then TEA (R = 1; where R = mole of TEA/mole of metal alkoxides) was added to create the B-site complex compound. In a separate vessel, to make the A-site precursor, lead acetate trihydrate was mixed with propyleneglycol and water in a 1:12:44 molar ratio. The solution was then heated to ∼70  C for dissolution of lead acetate trihydrate. This method was used to prepare all of Pb precursors. Finally, the A-site and

2.2. Thin-Film Fabrication Film deposition was carried out on the Pt (111)-coated {100} silicon substrates (Si/10,000Å-SiO2 /200Å-Ti/1500ÅPt) by spin-coating at 2800 rpm for 30 s. Following pyrolysis at 350–380  C for 3 min, additional layers were spin-coated to build up the desired thickness. The films were crystallized by the rapid insertion heating process at 700–750  C for 10 min. The final film thicknesses were about 330–350 nm. In order to investigate the effect of heat-treatment procedure on microstructure and related characteristics, three different conditions (F I, F II, F III) for the ferroelectric PMN-PT thin-film series and two different conditions (AF I, AF II) for the antiferroelectric/ferroelectric PZSTN thin-film series were selected, respectively. For the F I process, the film was singly crystallized at 730  C for 5 min after deposition of multilayer amorphous film. For the F II, the film was crystallized at 730  C for 5 min after heating the first layer at the same condition, while layer-by-layer heating at 730  C for 5 min was done in sequence for the F III process [39]. The antiferroelectric/ferroelectric PZSTN thin films were formed by crystallizing at 700  C for 10 min after heating each layer at 350  C (AF I) and after crystallizing each layer at 700  C for 5 min without pyrolysis at 350  C (AF II), respectively. In case of the thin films having the buffered structure, a total of seven layers was deposited with varying stacking sequences such as AFE layer/FE layer/AFE layer

Pb-Based Ferroelectric Nanomaterials

Electrical measurements were made through the film thickness using Pt top electrode 0.5 mm in diameter prepared by sputtering through a shadow mask. Electrical contact was made to the bottom electrode with a probe wire, whereas the top contact was made with a microprobe tip. Low field dielectric properties were determined by an impedance/gain phase analyzer (4194A, Hewlett-Packard, USA) with an oscillation of 10 mV and frequencies of 0.1 kHz to 100 kHz. The hysteresis loop of polarization (P) versus electric field (E) and the fatigue characteristics of each film were obtained with a modified computer-controlled Sawyer–Tower circuit using an RT66A ferroelectric tester (Radiant Tech., USA) in conjunction with an external function generator (19, Wavetek, USA) and digitizing oscilloscope (VC6025, Hitachi, Japan).

3. PHYSICAL PROPERTIES OF PMN AND PZT THIN FILMS

110

6

90

4

80

2

70

0

60

-2 0

200

400

600

800

Temperature (°C) 8

110

(b)

100

6

90

4

80

2

70

0

60

3.1. Solution Chemistry

-2 0

200

400

600

800

Temperature (°C) 8

110

Weight loss (%)

Since Budd et al. [34] first demonstrated the methoxyethanol route, the modification of this route has been done in many ways. Especially to control the hydrolysis and condensation reaction of parent alkoxides during solution preparation, a number of related sol–gel methods using several alkoxides stabilized by chelating agents, or using other alcohol solvent, have been proposed. Alkoxides of some metals, such as Zr, Ti, and Nb, react with -dikonates, diethanolamine, or triethanolamine to form chelate complexes [35, 39, 46]. These chelated complexes are less sensitive to moisture and result in stability to hydrolysis and precipitation during the solution preparation. Generally, thin films prepared by using solution stabilized by chelating agent show relatively high crystallization temperature and small grain size but little porous

8

(a)

100

(c)

100

6

90

4

80

2

70

0

60

0

200

400

600

800

-2

Temperature difference (°C)

2.3.2. Electrical Measurements

Temperature difference (°C)

The structure and crystallinity of the films were characterized by an X-ray diffraction (XRD; Rigaku Co., Japan) using CuK radiation. Typically step scans from 20 to 60 , at 0.04 increments and 4 s count time, were used. A differential thermal analysis (DTA) and thermogravimetry (TG) were performed on dried gel powders which were prepared by heating to 150  C for 24 h. DTA diagrams were obtained at a heating rate of 10  C/min and an airflow rate of 40 cm3 /min. Also organic residue of the film after coating each layer was analyzed by using FT-IR (1600, Perkin Elmer, USA), where the film was deposited on MgO(100) single crystal. The film morphology was observed by a scanning electron microscope (SEM; S4200, Hitachi, Japan).

Temperature difference (°C)

2.3.1. Spectroscopic Characterization

Weight loss (%)

2.3. Measurements and Characterization

microstructure because they contain more carboxyl ligand and require high thermodynamic driving force to decompose the organic compound. A good understanding of organic decomposition behavior of solution precursor and/or gel in chemical solution processing is essential to obtain more dense microstructure in chemical solution-derived thin films. DTA/TG diagrams of PMN, PZT, and 0.5PMN-0.5PZT are represented in Figure 1a, b, and c. The main exothermic peaks at 300–400  C are due to the combustion of the carboxylate compounds bound to the alkoxy groups in the precursors. This is also confirmed by a distinct weight loss in the TG data of Figure 1 [47]. It is noted that the temperature at which organic pyrolysis occurred is slightly different for the different compositions. In case of PMN gel powder, a large exothermic peak due to organic combustion appears at around 400  C, while for the PZT and 0.5PMN0.5PZT gel powders, it appeared at about 288  C and 360  C, respectively. The difference of organic pyrolysis temperature is mainly attributed to different chelating agents used in the preparation of each chemical solution as explained in the previous experimental section. This suggests that the pyrolysis behavior is related to the character of carboxylate groups in the PMN and PZT solution because different complexing agent was used to stabilize each solution. Tahan et al. [48] reported that DTA diagrams of dried gels were dependent on the mixed ratio of acetic acid to ethyleneglycol in (Ba, Sr)TiO3 solution processing. Thus, this difference is

Weight loss (%)

(AFE buffered FE), and FE layer/AFE layer/FE layer (FE buffered AFE), where PZT(50/50) and PZNST (40/15/2) thin films, and PZ and PZNST (40/3/2) thin films were applied as ferroelectric and antiferroelectric layers, respectively. The total thickness of the annealed films was about 330 nm.

437

Temperature (°C)

Figure 1. (a) DTA/TG curves of PMN solution-derived gel powder, (b) PZT solution-derived gel powder, and (c) 0.5PMN-0.5PZT solutionderived gel powder. Reprinted with permission from [47], J. H. Park et al., Ferroelectrics 260, 75 (2001). © 2001, Gordon and Breach.

Pb-Based Ferroelectric Nanomaterials

438

(e) Transmittance(arb.unit)

also explained by comparison of physical properties of the constituents used in this study. Table 1 represents the physical constants of typical solvent and various chelating agents [49]. As shown in Table 1, TEA has higher molecular weight and boiling point than those of acetylacetone or acetic acid. This reason is why organic pyrolysis temperature of the PMN solution is higher than that of the PZT solution. Two exothermic peaks of the Figure 1c solution also indicate the behavior of different chelating agents in 0.5PMN-0.5PZT solution, where peak temperatures were slightly different. In case of the modified chemical solution process, it may be more worthwhile to consider appropriate heat-treatment condition to eliminate organic content in the amorphous films and minimize the film cracking during the crystallization. Organic pyrolysis behavior was also investigated by FTIR spectroscopy analysis. The typical FT-IR transmittance spectra shown in Figure 2 ascertain that in case of 0.5PMN0.5PZT thin films, most organic groups were burned out in the temperature ranges of 350–380  C. Therefore, an intermediate pyrolysis at 380  C was performed in case of PMNcontaining thin films, while PZSTN thin films prepared by the conventional process based on 2-methoxyethanol were pyrolyzed at 350  C as explained in the experimental section. Figure 3 shows XRD patterns of the PMN, PZT, and 0.5PMN-0.5PZT thin films crystallized at optimum temperature that was experimentally determined in the range of 700–750  C. It can be found that all the films prepared by chemical solution have a single-phase perovskite structure within X-ray detection limit. In analyzing the lattice spacing of the materials, the thin-film spacings were slightly larger than in the bulk material. It may be due to the mechanical stress present in thin films caused by lattice or thermal expansion mismatch between the film and substrate, resulting in differences in lattice constant of bulk and corresponding thin-film materials. SEM photographs of the thin films corresponding to above compositions are shown in Figure 4. The PMN and PZT films showed uniform grain structures with an average grain size of about 400 nm and 150 nm, respectively. The 0.5PMN-0.5PZT films also showed uniform and dense microstructure with an average grain size

(d)

(c)

(b)

(a)

COOCOO(symmetric)

(asymmetric) 3500

3000

2500

2000

1500

1000

-1

Wavenumber (cm )

Figure 2. FT-IR transmittance spectra collected from thin layers by heating at different temperatures, (a) as deposited, (b) 200  C for 3 min, (c) 300  C for 3 min, (d) 350  C for 3 min, and (e) 380  C for 3 min.

around 200 nm. For the precursor modified with complexing agent or specific catalyst, it requires relatively higher temperature to decompose the organic compound and resulted in smaller grain size distribution or higher crystallization temperature. Similarly, Francis and Payne [50] reported that more uniform and small-grain-sized PMNT thin films were developed by adding benzoic acid as a catalyst. Figure 5 shows the frequency dependency of the dielectric constant and loss for various films by chemical solution. The room-temperature dielectric constants, measured at 1 kHz, of the PMN, PZT, 0.5PMN-0.5PZT, 0.9PMN0.1PZT, and 0.9PMN-0.1PT films, are summarized in Table 2

substrate

(111)

Chemical name Formula 2-methoxyethanol C3 H 8 O2 Propylenglycol C3 H 8 O2 Acetic acid C2 H 4 O2 Acetylacetone C5 H 8 O2 Water H2 O Triethanolamine C6 H15 NO3

Mw

bp



nD

7610

12443

09663

14028

7610

1882

1036

14324

6005

118

1053

1049

10012

14005

0976

14512

1801

10000

1000

1333

14919

3354

11242

14852

Note: Mw : molecular weight (g), bp : boiling point ( C) at 1 atm., : density (g/cm3  at 20 C, nD : refractive index at R.T. Source: Reprinted with permission from [49], C. R. Hammond, in “CRC Handbook of Chemistry and Physics” (D. R. Lide, Ed.). CRC Press, Boca Raton, FL, 1995. © 1995, Chemical Rubber Company Press.

Intensity(arb. unit)

Table 1. Physical properties of typical solvents and chelating agents. substrate

Pt3Ti

(110)

(c)

(b)

(a) 20

25

30

35

40

45

50

55

60

2θ (degree)

Figure 3. XRD patterns of solution-derived ferroelectric thin films. (a) PMN, (b) PZT, and (c) 0.5PMN-0.5PZT thin film. Reprinted with permission from [47], J. H. Park et al., Ferroelectrics 260, 75 (2001). © 2001, Gordon and Breach.

Pb-Based Ferroelectric Nanomaterials

(a)

439 Table 2. Summary of the dielectric and hysteresis properties of solution-derived Pb-based ferroelectric thin films by chemical solution deposition.

(b)

(c)

Thin films

r1 kHz

(%)a

tan

Pr (C/cm2 )

Ec (kV/cm)

PMN 0.5PMN-0.5PZT 0.1PMN-0.9PZT PZT 0.9PMN-0.1PT

2600 1994 1420 926 2750

131 135 71 68 127

003 003 003 0027 0031

13 143 18 113 58

45 348 40 485 11

a Degree of dispersion for the dielectric constant with frequency ( ) was calculated from the equation  r1 kHz − r100 kHz / r1 kHz × 100 %. Source: Reprinted with permission from [47], J. H. Park et al., Ferroelectrics 260, 75 (2001). © 2001, Gordon and Breach.

3.2. Nucleation and Grain Growth Figure 4. SEM photographs of chemical solution-derived ferroelectric thin films. (a) PMN, (b) PZT, and (c) 0.5PMN-0.5PZT thin film. Reprinted with permission from [47], J. H. Park et al., Ferroelectrics 260, 75 (2001). © 2001, Gordon and Breach.

[47, 51, 52]. The dielectric properties of the Pb-based thin films were comparable or superior to those of the films with same composition prepared by conventional sol–gel process as reported previously [50, 53, 54]. This confirms that the chemical solution method modified by the chelating agent used in this study can be applied to prepare a reliable ferroelectric thin film with a variety of lead-based compositions.

PZT 0.1PMN-0.9PZT 0.5PMN-0.5PZT PMN 0.9PMN-0.1PT

(111)

2500

2000

Perovskite Substrate

1k

10k

100k

Pt CuKβ

(100)

(c) (211)

500

(110)

1000

(200)

1500

Intensity (arb. unit)

Dielectric constant (K)

3000

Figure 6 shows typical XRD patterns for 0.5PMN0.5PT(50PMNT) thin films prepared according to the multilayer film formation and heating process as explained in experimental part. These diffraction patterns showed that they were developed highly (100)-oriented ones (F I film) (Fig. 3a) while the highly (111)-oriented films were obtained for F I and F II films (Fig. 3b and c). In regard to both (111)- and (100)-oriented thin films, Liu and Phule [55] reported that the (111) nuclei grew and a strong (111) texture developed in sol–gel-derived PZT film, when the film that had a perovskite seed layer formed in-situ was subjected to a higher-temperature heat treatment. Tani et al. [56] suggested that (100) PLZT texture was developed, when the film was deposited onto platinized Si substrate which was free of any intermetallic phase between Pt and Ti. Further, they also concluded that this (100) texture represented minimum surface energy. Figure 7 shows the SEM photographs of the 50PMNT films with different multilayer film formation conditions. The thin film deposited by F I process consisted of small and large grains with grain size distribution of

(b)

Frequency (Hz)

(a) tanδ

0.16 0.14 0.12 0.10 0.08 0.06 0.04 0.02

Figure 5. Frequency dependencies of dielectric constant and tan for the various ferroelectric thin films by chemical solution deposition.

20

25

30

35

40

45

50

55

60

2θ (degree)

Figure 6. XRD patterns of solution-derived 0.5PMN-0.5PT thin films deposited on Pt-passivated Si by (a) F I, (b) F II, and (c) F III processes. Reprinted with permission from [41], J. H. Park et al., J. Am. Ceram. Soc. 82, 2116 (1999). © 1999, American Ceramic Society.

Pb-Based Ferroelectric Nanomaterials

440 (b)

(a)

(a)

0.5 µm

Figure 8. SEM photographs of the (a) 10 mol% PMN modified PZT(50/50) and (b) Zr/Ti modified PZT (80/20).

(d)

(c)

0.5 µm

0.5 µm

0.5 µm

(b)

4. ELECTRICAL FATIGUE OF PZSTN THIN FILMS 4.1. Grain Size Effect 0.5 µm

Figure 7. SEM photographs of the 0.5PMN-0.5PT thin films by (a) F I, (b) F II, (c) F III processes, and (d) cross-sectional view of (b). Reprinted with permission from [41], J. H. Park et al., J. Am. Ceram. Soc. 82, 2116 (1999). © 1999, American Ceramic Society.

about 50 to 200 nm. On the other hand, the grain size of the film by F II process became smaller. For the film by the F III process shown in Figure 4c, its grain size reduced effectively up to average grain size under 20 nm. It means that introduction of a perovskite seed layer before a crystallization of amorphous thin film could facilitate the microstructural change of the films into more small and uniform grain size distribution. It is possibly ascribed to the presence of more nucleation sites between Pt substrate and the film or in the film, resulting in smaller grain size distribution [40, 57, 58]. The dielectric and ferroelectric properties of the films with different multilayer film formation conditions are summarized in Table 3. The dielectric constant of the F I film was slightly larger compared to that of F II or F III films heat-treated at 730  C. These results imply that the degree of orientation and microstructure with different grain sizes have effects on the weak-field dielectric properties. A similar correlation between the orientation and grain size distribution has been also noticed by Aoki et al. [59] on sol–gelderived PZT films, even when no pure perovskite phase was observed. Figure 8 indicates that the microstructure of PZT thin film effectively changed through the addition of relaxor compound and an increase of zirconium to titanium ratio, confirming that nanocrystallization of thin film can be controlled by compositional modification as well as crystallization condition.

PZSTN compositions are also known to be attractive material for actuator and transducer applications due to relatively easier introduction of antiferroelectric and ferroelectric phase, leading to large displacement [60–63]. The antiferroelectric (40/3/2) and ferroelectric (40/15/2) PZSTN thin films were prepared by different heat treatments (AF I, AF II, AF III) and their SEM photographs are shown in Figure 9. The films prepared by the AF I process consisted of the large grains for both films (Fig. 9a, b), while, in the films prepared by the AF II, the nanocrystalline microstructure of very fine grains under 100 nm was obtained, where the 40/15/2 film showed a little larger grains than 40/15/2 possibly due to the higher content of ferroelectric (PZT) phase. The antiferroelectric film (40/3/2) has small values of switchable polarization compared to those of the ferroelectric film (40/15/2). The saturated polarization (Ps), corresponding to the difference between switched polarization (P*) and the half value of switchable polarization ((P∗ r-P∧ r)/2), has been introduced [32, 40]. The change of normalized Ps in the PZSTN thin films is shown in Figure 10. The degradation of polarization in the ferroelectric composition (40/15/2) was steeper than that in the (a)

(b)

(c)

(d)

Table 3. Comparison of dielectric properties of the 0.5PMN-0.5PT thin films deposited with various annealing conditions. Annealing condition FI F II F III

Dielectric constant (at 1 kHz)

tan (at 1 kHz)

Preferential orientation

1690 1378 1413

0.02 0.03 0.03

100 111 111

Source: Reprinted with permission from [41], J. H. Park et al., J. Am. Ceram. Soc. 82, 2116 (1999). © 1999, American Ceramic Society.

Figure 9. SEM photographs of the PZSTN thin films with antiferroelectric (40/3/2) composition (a) by the AF I and (c) by the AF II process; and ferroelectric (40/15/2) composition (b) by the AF I process and (d) by the AF II process.

Pb-Based Ferroelectric Nanomaterials (b) 1.0

1.0

0.9

0.9

Normalized Ps

Normalized Ps

(a)

441

0.8 0.7 AF I process AF II process

0.6

(b)

0.8 0.7 AF I process AF II process

0.6 0.5

0.5 10

(a)

2

10

4

10

6

Cycles

10

8

10

10

10

2

10

4

10

6

10

8

10

10

Cycles 4 µm

Figure 10. Comparison of the fatigue curves between the films prepared by the AF I process and the AF II process for (a) antiferroelectric (40/3/2) and (b) antiferroelectric (40/15/2) compositions.

antiferroelectric composition (40/3/2). It was considered that the antiferroelectrics contained mainly 180 domains which had smaller internal stress during switching cycles compared to 90 domains as reported elsewhere [64]. The following things seem to be more interesting. The PZSTN thin films prepared by the AF I process (the large and irregular grain size) showed more degradation of polarization than the films prepared by the AF II process showing nanocrystalline structure with regular grain size for both compositions. The steeper degradation of polarization in the films prepared by the AF I process could be mainly attributed to the larger grains and irregular microstructure. The different distributions of the applied electric field and internal stress between the large perovskite grains and the grain boundaries (or the different phases between grains) during repetitive switching should generate the electrical and mechanical defects especially at the interfaces of the two different phases. The uniformly distributed nanocrystalline structure could release the stress coming from the above reason more than the microstructure consisting of the large grains. Similar phenomena will be found in the results shown in the next section. The phase images of electrostatic charge were taken using a commercial scanning force microscope. Initially, a dc voltage of −10 V was applied while a conductive tip scanned over the desired area. Subsequently, the phase images were obtained under applying an ac voltage of 1 V and a frequency of 17 kHz. Details of instrumentation and mechanisms for the imaging have already been published [65]. Figure 11 shows phase images of electrostatic charge of the 40/3/2 and 40/15/2 compositions. After being scanned with a conductive tip held under an applied dc voltage of −10 V, the thin film having 40/3/2 composition showed a quite different phase image compared with the film having 40/15/2 composition. In the 40/3/2 film [Fig. 11a], the bright regions of the phase image result from the polarization directions toward the bottom electrode while the dark regions result from the polarization directions toward the top electrode. The bright and dark regions seemed to consist of several nanocrystalline grains. This behavior implies that the very small nano-sized domains have their antiferroelectric properties. It was confirmed that the antiferroelectric 40/3/2 composition was switched by 180 . On the other hand, the phase image of the 40/15/2 film [Fig. 11b] showed no contrast in the whole area. This configuration also means that the full

Figure 11. Phase images of electrostatic charge of the PZSTN thin films with (a) antiferroelectric (40/3/2) and (b) ferroelectric (40/15/2) compositions after applying −10 V. Reprinted with permission from [68], J. H. Jang et al., Appl. Phys. Lett. 73, 1823 (1998). © 1998, American Institute of Physics.

nano domains of the 40/15/2 have switched to the unipolar direction and switched by 90 .

4.2. Buffer-Layer Effect In addition to the effect of the overall grain size of the PZSTN thin films on the electrical properties, we have modified the grain size of very thin buffer layers existing on the interface between the electrode and the main body. The buffer layers have the PZT and PZSTN system but utilized different compositions to the main dielectric films. The PZ buffer layer on PZT body has grain size less than 20 nm as shown in Figure 12a. The PZT buffer layer on PZ body has two different grain sizes of about 20 nm, same as PZ buffer, and 60–70 nm as shown in Figure 12b. The change of P*r-P∧ r as a function of switching cycles of PZT, PZ buffered PZT, and PZT buffered PZ is shown in Figure 13 [64]. P*r is the switched remanent polarization between two opposite polarity pulses, and P∧ r is the nonswitched remanent polarization between two same polarity pulses. The difference between P*r and P∧ r denotes the switchable remanent polarization, an important variable for nonvolatile memory application. The PZT films showed a significant drop in polarization after 105 cycles. The PZ buffered PZT did not show any drop of polarization up to 109 cycles with sufficient remanent polarization of about 5 C/cm2 . The virgin state of the PZ buffered PZT has smaller remanent polarization than that of the PZT. However, the relative values were reversed after 2 × 106 cycles. This implies that the nanocrystalline antiferroelectric

(a)

(b)

Figure 12. SEM photographs of (a) AFE-PZ buffer on FE-PZT and (b) FE-PZT buffer on AFE-PZ thin films.

Pb-Based Ferroelectric Nanomaterials

442

FE AFE buffered FE FE buffered AFE

Polarization (µC/cm2)

15

10

5

0 102

103

104

105

106

107

108

109

Cycles

Figure 13. Switchable polarization (P*r-P∧ r) based on ferroelectric PZT(FE) and antiferroelectric PZ(AFE) thin films. Reprinted with permission from [64], J. H. Jang and K. H. Yoon, Appl. Phys. Lett. 75, 130 (1999). © 1999, American Institute of Physics.

layer should act as a barrier to degradation of polarization. Reduced stress related to 180 domains during switching would suppress the formation of electrical and mechanical defects, which cause the generation of the oxygen vacancies [42, 43]. The near zero fatigue in the PZ buffered PZT compared with that of the PZT films was attributed to the relatively low stress and the resulting low number of oxygen vacancies induced at the ferroelectric or antiferroelectric-Pt electrode interface [66, 67]. With the antiferroelectric phase, the nanocrystalline microstructure played an important role to reduce the internal stress. It was confirmed by the fact that PZT buffered PZ also showed very good fatigue endurance as shown in Figure 13, although it had an insufficient remanent polarization value for FRAM application. The saturated polarization (Ps) of AF, AF buffered FE, FE buffered AF, and FE is shown in Figure 14 [67]. All films except FE showed near zero fatigue after 109 cycles.

Normalized Ps

1.0

0.9

AF AF buffered FE 0.8

(1)

dQ = V × A × T /R

where V is the voltage applied across the film, A is the capacitor area, T is the time increment, and R is the resistance for polarity of the applied voltage. In the hysteresis loop after 105 cycles, the remanent polarization increased compared with the initial state, and such behavior agreed with the P*r-P∧ r data in Figure 13. The increase of remanent polarization near 105 cycles should be related to the release of the suppressed polarization originating from the difference in lattice parameters between PZT and PZ. The (a)

(b)

(c)

20 10 0 -10 -20 2 1 0 -1 -2 -400 -200

0

200

Field (kV/cm)

400 -400 -200 0

200 400 -400 -200

Field (kV/cm)

0

200 400

Field (kV/cm)

FE buffered AF FE

102

The FE and AF buffered FE showed sufficient remanent polarization, while FE buffered AF and AF showed insufficient remanent polarization. The AF buffered FE showed excellent fatigue properties maintaining more than 90% of initial polarization values after 109 cycles. FE showed the stiffest degradation of polarization. FE buffered AF showed better fatigue endurance than FE. The near zero fatigue in the AF buffered FE was attributed to the fact that the nanocrystalline microstructure having antiferroelectric phase should act as a barrier to degradation of polarization. The relatively lower stress and the resulting smaller number of oxygen vacancies should be induced at the nanocrystalline buffer-Pt electrode interface than at the large grains-Pt electrode interface. Figure 15 shows the hysteresis loops and the leakage current component of the PZ buffered PZT film with different switching cycles [64, 67]. The hysteresis loops and leakage current component of the corresponding hysteresis loop were measured using the same sample, using “Analysis” software in RT66A ferroelectric tester. During the voltage step, the change in charge (dQ) due to the leakage current can be estimated from the measured resistance by the following equation:

Polarization (µC/cm2) Polarization (µC/cm2)

20

103

104

before after 105 cycles 105

106

107

108

after 107 cycles after 109 cycles

109

Cycles

Figure 14. Switchable polarization based on ferroelectric PZSTN(FE) and antiferroelectric PZSTN(AFE) thin films. Reprinted with permission from [67], J. H. Jang and K. H. Yoon, Thin Solid Films 401, 67 (2001). © 2001, Elsevier Science.

Figure 15. The change of P-E hysteresis loops and leakage current component of (a) the PZ buffered PZT thin film, (b) the FE, and (c) the AF buffered FE during fatigue. Reprinted with permission from [64], J. H. Jang and K. H. Yoon, Appl. Phys. Lett. 75, 130 (1999). © 1999, American Institute of Physics; and from [67], J. H. Jang and K. H. Yoon, Thin Solid Films 401, 67 (2001). © 2001, Elsevier Science.

Pb-Based Ferroelectric Nanomaterials

release of polarization occurred asymmetrically. This behavior was attributed to the different natures of crystallization between the two PZ layers, because one PZ layer was crystallized on the Pt bottom electrode and the other PZ layer was crystallized on PZT layers during the deposition process of the films. As well as near 105 cycles, the leakage current component of the corresponding hysteresis loop (Fig. 15a) increased negligibly after 109 cycles. This result supports the release of the suppressed polarization. The resistivity measurement also indicated that the resistivity of the PZ1/PZT5/PZ1 film did not change after 109 cycles. In the hysteresis loop of FE (Fig. 15b), the increase of remanent polarization in hysteresis loop of FE was found after 107 cycles. The increase in FE should come from the leakage current component of the corresponding hysteresis loop increased after 107 cycles. The reason for leakage in fatigued FE could be explained by the large displacement of the grains under the applied field and the large stress during the 90 domain switching. The AF buffered FE showed nearly no increase of remanent polarization and leakage current component after 109 cycles as shown in Figure 15c. There should be a small amount of defect dipoles due to nanocrystalline antiferroelectric buffer. The negligible leakage current component could be explained by the reason that the antiferroelectric buffer has very small grain size and internal stress during 180 switching acts as a barrier to fatigue between the electrode and the ferroelectric layers. Therefore, the antiferroelectric buffer could be resistant to mechanical stress and resulting microcracking or production of electrical defects. The AF buffered FE showed somewhat different fatigue properties from the PZ buffered PZT. In the PZ buffered PZT, there was the release of the suppressed polarization originating from the mismatches in lattice parameters and microstructures between PZT and PZ, while there was no significant change in AF buffered FE during 109 cycles of ±10 V due to the similar lattice parameters and microstructures [64].

5. CONCLUSIONS We have outlined the chemical synthesis of Pb-based nanocrystalline perovskite thin films and their effects on the electrical properties such as dielectric constant, hysteresis loop, and especially fatigue character. For Pb-based perovskites, the modification of metal alkoxide by a chemical solution processing gave rise to the formation of a stable complex against hydrolysis, and eventually enhanced reliability and uniformity of the films. Surface microstructural characteristics resulting from the crystallization of as-prepared thin films were found to depend strongly on the various multilayer film formations as well as solution chemistry. Moreover, with proper control of multilayer thin-film formation procedure, it may be possible to control its related dielectric properties as well as preferred orientation of the ferroelectric thin films. For antiferroelectric and ferroelectric PZSTN thin films, nanocrystalline microstructure was obtained with various heat treatments and compositional change, and the modified multilayer processes improved the electric fatigue

443 endurance. From the scanning force microscopy, it was confirmed that the very small nano domains have their ferroelectric and antiferroelectric properties. The nanocrystalline antiferroelectric buffer has improved the electric fatigue endurance of the PZT-based ferroelectric thin films prominently. The P*r-P∧ r of the PZ buffered PZT films and AF buffered FE (PZSTN) showed near zero degradation of polarization after 109 cycles on the Pt electrode. The nanocrystalline antiferroelectric PZ and AF buffers acted as barriers to fatigue due to their 180 domain switching and stress release in the nanocrystalline microstructure. It is expected to be applied to other popular thin-film processes for nonvolatile memory applications and other fields, such as the actuator which needs long-term reliability.

GLOSSARY Antiferroelectric (AFE) There is no net polarization of a cell in the absence of an electric field, but the cell is polarized when a field is applied. An antiferroelectric material will be strongly repelled by an electric field. The dipole moments in an antiferroelectric are arranged with an equal number pointing in each direction. Electric fatigue The loss in switchable polarization at a fixed drive voltage as a function of the continuous switching of a ferroelectric capacitor. Ferroelectric (FE) A kind of electric phase in crystal where the centers of the positive and negative charges do not coincide even without the application of external electric field. In this case, spontaneous polarization exists in the crystal. And the polarization of the dielectric can be reversed by an electric field. Microelectromechanical system (MEMS) Devices that have a characteristic length of less than 1 mm but more than 1 m, that combine electrical and mechanical components that are fabricated using integrated circuit batch-processing technologies. Nanoelectromechanical system (NEMS) The NEMS devices that have the dimension of nanometer range. Sol–gel process Colloidal route used to synthesize chemicals with an intermediate stage including a sol/or a gel state. (Sol–gel is restricted to the gels synthesized from alkoxides).

REFERENCES 1. J. F. Scott and C. A. Paz de Araujo, Science 246, 1400 (1989). 2. D. Dimos, S. J. Lockwood, and R. W. Schwartz, IEEE Trans. Comp. Pack. Man. Tech. A18, 174 (1995). 3. V. Bobnar, Z. Kutnjak, A. Levstik, J. Holc, M. Kosec, T. Hauke, R. Steinhausen, and H. Beige, J. Appl. Phys. 85, 622 (1999). 4. T. Tani, J. Li, D. Viehland, and D. A. Payne, J. Appl. Phys. 75, 3017 (1994). 5. H. D. Chen, K. R. Udayakumar, C. J. Gaskey, and L. E. Cross, Appl. Phys. Lett. 67, 3411 (1995). 6. V. Nagarajan, S. P. Alpay, C. S. Ganpule, B. K. Nagaraj, S. Aggarwal, E. D. Williams, A. L. Roytburd, and R. Ramesh, Appl. Phys. Lett. 77, 438 (2000). 7. T. Haccart, E. Cattan, D. Remiens, S. Hiboux, and P. Muralt, Appl. Phys. Lett. 76, 3292 (2000).

444 8. A. M. Flynn, L. S. Tavrow, S. F. Bart, R. A. Brooks, D. J. Ehrlich, K. R. Udayakumar, and L. E. Cross, J. Microelectromechan. Syst. 1, 44 (1992). 9. K. R. Udayakumar, J. Chen, A. M. Flynn, S. F. Bart, L. S. Tavrow, D. J. Ehrlich, L. E. Cross, and R. A. Brooks, Ferroelectrics 160, 347 (1994). 10. G. H. Berrstein, H. V. Goodson, and G. L. Snider, in “The MEMS Handbook” (M. Gad-el-Hak, Ed.). CRC Press, Boca Raton, FL, 2002. 11. R. P. Feynman, J. Microelectromechan. Syst. 1, 60 (1992). 12. F. Cerrina, Proc. IEEE 84, 644 (1997). 13. S. Matsui, Proc. IEEE 84, 629 (1997). 14. M. J. Mayo, D. C. Hague, and D. J. Chen, Mater. Sci. Eng. A166, 145 (1993). 15. K. P. Kumar, K. Kelzer, A. J. Burggraaf, T. Okuba, H. Nagamoto, and S. Morooka, Nature 358, 48 (1992). 16. W. H. Rhodes, J. Am. Ceram. Soc. 64, 19 (1981). 17. K. I. Choy, in “Handbook of Nanostructured Materials and Nanotechnology” (H. S. Nalwa, Ed.), Vol. 1, p. 533. Academic Press, San Diego, 2000. 18. Q. Yitai, in “Handbook of Nanostructured Materials and Nanotechnology” (H. S. Nalwa, Ed.), Vol. 1, p. 459. Academic Press, San Diego, 2000. 19. Q. W. Chen, Y. T. Qian, Z. Y. Chen, W. B. Wu, Z. W. Chen, G. E. Zhou, and Y. H. Zhang, Appl. Phys. Lett. 66, 1 (1995). 20. Q. W. Chen, Y. T. Qian, Z. Y. Chen, G. E. Zhou, and Y. H. Zhang, Mater. Lett. 22, 93 (1995). 21. Q. W. Chen and Y. T. Qian, Thin Solid Films 264, 25 (1995). 22. Q. W. Chen, Y. T. Qian, Z. Y. Chen, Y. Xie, G. E. Zhou, and Y. H. Zhang, Mater. Lett. 24, 85 (1995). 23. Q. W. Chen, Y. T. Qian, Z. Y. Chen, L. Shi, X. G. Li, G. E. Zhou, and Y. H. Zhang, Thin Solid Films 272, 1 (1996). 24. Y. Xie, W. Z. Wang, Y. T. Qian, L. Yang, and Z. Chen, J. Crystal Growth 167, 656 (1996). 25. W. J. Desisto, Y. T. Qian, C. Hanni Gan, J. O. Edward, R. Kershaw, K. Dwight, and A. World, Mater. Res. Bull. 25, 183 (1990). 26. Q. W. Chen, X. G. Li, Y. T. Qian, J. S. Zhu, G. E. Zhou, W. P. Zhang, and Y. H. Zhang, Appl. Phys. Lett. 68, 1 (1995). 27. Y. T. Qian, Y. Xie, Z. Chen, J. Lu, and J. Zhu, J. Chem. Phys. 8, 549 (1995). 28. W. S. Hu, Z. G. Liu, Z. C. Wu, and D. Feng, Mater. Lett. 28, 369 (1996). 29. T. C. Chou, D. Adamson, J. Mardinly, and T. G. Nieh, Thin Solid Films 205, 131 (1991). 30. X. Mei, M. Tao, H. Tan, Y. Han, and W. Tao, Mater. Res. Soc. Symp. Proc. 286, 179 (1993). 31. T. Sugino, K. Tanioka, S. Kawasaki, and J. Shirafuji, Diamond Related Mater. 7, 632 (1998). 32. T. Atsuki, N. Soyama, G. Sasaki, T. Yonezawa, K. Ogi, K. Sameshima, K. Hoshiba, Y. Nakao, and A. Kamisawa, Jpn. J. Appl. Phys. 33, 5196 (1994). 33. M. N. Kamalasanan, N. D. Kumar, and S. Chandra, J. Appl. Phys. 76, 4603 (1994). 34. K. D. Budd, S. K. Dey, and D. A. Payne, Brit. Ceram. Soc. Proc. 36, 107 (1985). 35. G. Yi, Z. Wu, and M. Sayer, J. Appl. Phys. 64, 2717 (1988). 36. Y. L. Tu and S. J. Milne, J. Mater. Res. 10, 3222 (1995).

Pb-Based Ferroelectric Nanomaterials 37. Y. Takahashi, Y. Matsuoka, K. Yamaguchi, M. Matski, and K. Kobayashi, J. Mater. Sci. 25, 3960 (1990). 38. K. H. Yoon, J. H. Park, and D. H. Kang, J. Am. Ceram. Soc. 78, 2267 (1995). 39. J. H. Park, K. H. Yoon, and D. H. Kang, J. Am. Ceram. Soc. 82, 2683 (1999). 40. K. H. Yoon, J. H. Park, and J. H. Jang, J. Mater. Res. 14, 2933 (1999). 41. J. H. Park, K. H. Yoon, and D. H. Kang, J. Am. Ceram. Soc. 82, 2116 (1999). 42. J. H. Jang and K. H. Yoon, Ferroelectrics 225, 193 (1999). 43. Q. Y. Jiang, E. C. Subbarao, and L. E. Cross, J. Appl. Phys. 75, 7433 (1995). 44. S. S. Sengupta, D. Roberts, J.-F. Li, M. C. Kim, and D. A. Payne, J. Appl. Phys. 78, 1171 (1995). 45. J. H. Park, K. H. Yoon, and D. H. Kang, Thin Solid Films 396, 84 (2001). 46. Y. Takahashi and Y. Wada, J. Electrochem. Soc. 137, 267 (1990). 47. J. H. Park, K. H. Yoon, and D. H. Kang, Ferroelectrics 260, 75 (2001). 48. D. M. Tahan, A. Safari, and L. C. Klein, J. Am. Ceram. Soc. 79, 1593 (1996). 49. C. R. Hammond, in “CRC Handbook of Chemistry and Physics” (D. R. Lide, Ed.). CRC Press, Boca Raton, FL, 1995. 50. L. F. Francis and D. A. Payne, J. Am. Ceram. Soc. 74, 3000 (1991). 51. J. H. Park, K. H. Yoon, D. H. Kang, and E. S. Kim, Ferroelectrics 258, 303 (2001). 52. J. H. Park, K. H. Yoon, D. H. Kang, and J. H. Park, Mater. Chem. Phys. 79, 151 (2003). 53. K. Okuwada, M. Imai, and K. Kakuno, Jpn. J. Appl. Phys. 28, L1271 (1989). 54. K. R. Udayakumar, J. Chen, P. J. Schuele, L. E. Cross, V. Kumar, and S. B. Krupanidhi, Appl. Phys. Lett. 60, 1187 (1992). 55. Y. Liu and P. P. Phule, J. Am. Ceram. Soc. 79, 495 (1996). 56. T. Tani, Z. Xu, and D. A. Payne, Ferroelectric Thin Films III 310, 269 (1993). 57. K. C. Chen and J. D. Mackenzie, Mater. Res. Soc. Symp. Proc. 180, 663 (1990). 58. C. H. Peng and S. B. Desu, Mater. Res. Soc. Symp. Proc. 243, 335 (1992). 59. K. Aoki, Y. Fukuda, K. Numuta, and A. Nishimura, Jpn. J. Appl. Phys. 32, 5155 (1994). 60. K. Uchino, Jpn. J. Appl. Phys. 24, 460 (1985). 61. C. Zhiming. L. Jingyu, and W. Yongling, Ferroelectrics 101, 225 (1990). 62. P. Yang and D. A. Payne, J. Appl. Phys. 71, 1361 (1992). 63. D. Viehland, D. Forst, Z. Xu, and J.-F. Li, J. Am. Ceram. Soc. 78, 2101 (1995). 64. J. H. Jang and K. H. Yoon, Appl. Phys. Lett. 75, 130 (1999). 65. J. Lee, S. Esayan, A. Safari, and R. Ramesh, Appl. Phys. Lett. 65, 254 (1994). 66. J. H. Jang, K. H. Yoon, and K. Y. Oh, Mater. Res. Bull. 35, 393 (2000). 67. J. H. Jang and K. H. Yoon, Thin Solid Films 401, 67 (2001). 68. J. H. Jang, K. H. Yoon, and H. J. Shin, Appl. Phys. Lett. 73, 1823 (1998).